Limiting Diffusion Coefficients for Ions and Nonelectrolytes in ...

10
Vol:.(1234567890) Journal of Solution Chemistry (2019) 48:748–757 https://doi.org/10.1007/s10953-019-00884-0 1 3 Limiting Diffusion Coefficients for Ions and Nonelectrolytes in Solvents Water, Methanol, Ethanol, Propan‑1‑ol, Butan‑1‑ol, Octan‑1‑ol, Propanone and Acetonitrile at 298 K, Analyzed Using Abraham Descriptors Michael H. Abraham 1  · William E. Acree Jr. 2 Received: 15 November 2018 / Accepted: 23 February 2019 / Published online: 21 May 2019 © The Author(s) 2019 Abstract We have used literature data on the conductivity of single ions to obtain limiting diffusion coefficients (D o ) of ions in water, alcohols, propanone and acetonitrile. We then used litera- ture data on limiting diffusion coefficients of nonelectrolytes in order to set up linear free energy relationships that combine data for ions and nonelectrolytes, as log 10 (D o ) in the same equation, all values being at 298 K. These equations are for solvents water (N = 377, SD = 0.0474), methanol (N = 96, SD = 0.0332), ethanol (N = 96, SD = 0.0666), propan-1-ol (N = 71, SD = 0.0487), butan-1-ol (N = 53, SD = 0.0600), octan-1-ol (N = 61, SD = 0.0684), propanone (N = 86, SD = 0.0366) and acetonitrile (N = 74, SD = 0.0438) where N is the number of data points, that is ions plus nonelectrolytes, and SD is the standard deviation in log 10 (D o ). It is shown that solute hydrogen bond acidity, solute hydrogen bond basicity and solute volume all lower the diffusion constants of ions and nonelectrolytes in the various solvents studied. Keywords Ions · Nonelectrolytes · Diffusion coefficients · Linear free energy relationship · Hydrogen bond acidity · Hydrogen bond basicity 1 Introduction Diffusion coefficients of solutes in water and in organic solvents play a very important role in chemical engineering [1], environmental chemistry [2] and in biological chemistry [3]. Therefore it is not surprising that there have been numerous studies on the correlation and prediction of limiting diffusion coefficients, mostly diffusion coefficients in water but also Electronic supplementary material The online version of this article (https://doi.org/10.1007/s1095 3-019-00884-0) contains supplementary material, which is available to authorized users. * Michael H. Abraham [email protected] 1 Department of Chemistry, University College London, 20 Gordon Street, London WC1H OAJ, UK 2 Department of Chemistry, University of North Texas, 1155 Union Circle Drive #305070, Denton, TX 76203, USA

Transcript of Limiting Diffusion Coefficients for Ions and Nonelectrolytes in ...

Vol:.(1234567890)

Journal of Solution Chemistry (2019) 48:748–757https://doi.org/10.1007/s10953-019-00884-0

1 3

Limiting Diffusion Coefficients for Ions and Nonelectrolytes in Solvents Water, Methanol, Ethanol, Propan‑1‑ol, Butan‑1‑ol, Octan‑1‑ol, Propanone and Acetonitrile at 298 K, Analyzed Using Abraham Descriptors

Michael H. Abraham1  · William E. Acree Jr.2

Received: 15 November 2018 / Accepted: 23 February 2019 / Published online: 21 May 2019 © The Author(s) 2019

AbstractWe have used literature data on the conductivity of single ions to obtain limiting diffusion coefficients (Do) of ions in water, alcohols, propanone and acetonitrile. We then used litera-ture data on limiting diffusion coefficients of nonelectrolytes in order to set up linear free energy relationships that combine data for ions and nonelectrolytes, as log10 (Do) in the same equation, all values being at 298 K. These equations are for solvents water (N = 377, SD = 0.0474), methanol (N = 96, SD = 0.0332), ethanol (N = 96, SD = 0.0666), propan-1-ol (N = 71, SD = 0.0487), butan-1-ol (N = 53, SD = 0.0600), octan-1-ol (N = 61, SD = 0.0684), propanone (N = 86, SD = 0.0366) and acetonitrile (N = 74, SD = 0.0438) where N is the number of data points, that is ions plus nonelectrolytes, and SD is the standard deviation in log10 (Do). It is shown that solute hydrogen bond acidity, solute hydrogen bond basicity and solute volume all lower the diffusion constants of ions and nonelectrolytes in the various solvents studied.

Keywords Ions · Nonelectrolytes · Diffusion coefficients · Linear free energy relationship · Hydrogen bond acidity · Hydrogen bond basicity

1 Introduction

Diffusion coefficients of solutes in water and in organic solvents play a very important role in chemical engineering [1], environmental chemistry [2] and in biological chemistry [3]. Therefore it is not surprising that there have been numerous studies on the correlation and prediction of limiting diffusion coefficients, mostly diffusion coefficients in water but also

Electronic supplementary material The online version of this article (https ://doi.org/10.1007/s1095 3-019-00884 -0) contains supplementary material, which is available to authorized users.

* Michael H. Abraham [email protected]

1 Department of Chemistry, University College London, 20 Gordon Street, London WC1H OAJ, UK2 Department of Chemistry, University of North Texas, 1155 Union Circle Drive #305070, Denton,

TX 76203, USA

749Journal of Solution Chemistry (2019) 48:748–757

1 3

in organic solvents [4]. Many of the older studies on diffusion coefficients have involved a restricted number of solutes, but recently there have been a number of studies which dealt with much larger numbers of solutes. Gharagheizi and Sattari [5] used a data base of diffusion coefficients in water for 300 solutes. They calculated 1664 descriptors for each solute from the Dragon software [6], and reduced these to the best six descriptors to use in a linear equa-tion. Garagheizi et al. [7] later used a huge data base of 4852 solutes to develop a predictive method based on 148 functional groups. Unfortunately, although they give the predicted dif-fusion coefficients for all 4852 solutes, the authors do not give any actual experimental values. Garagheizi [8] later used the 4852 data base to obtain a linear three parameter equation using descriptors calculated from the Dragon software program.

Use of extensive data bases has not been confined to diffusion coefficients in water. Magal-haes et al. [9] studied 3463 values of diffusion coefficients in water and in organic solvents. However, since 18 solvents were examined, the number of solutes per solvent is not that large. Li et al. [10] used a linear free energy relationship based on the Abraham equation [11], to correlate 743 diffusion coefficients in 25 solvents; again, although a large number of data points was used, the number per solvent is still rather limited.

All these recent studies involve only nonelectrolytes as solutes. Ions and ionic species are excluded. Indeed it seems impossible for methods based on the calculation of Dragon descrip-tors or on functional group analysis to be extended to ionic molecules. The only study of limit-ing diffusion coefficients that included ions and ionic molecules has been our own work on solutes in water [12]. It is the aim of the present work to enlarge the data base of ionic and nonelectrolyte solutes in water, and to extend our analysis to as many organic solvents for which data on nonelectrolytes, ions and ionic species are available.

2 Methods

Our general method is based on the linear free energy equation, LFER, Eq. 1.

The dependent variable is log10 (Do), where Do is the diffusion coefficient at infinite dilution at 298 K for a series of solutes in a given solvent. The independent variables in Eq. 1 (E, S, A, B, V, J+ and J−) are properties or descriptors of the solutes and the coef-ficients (c, e, s, a, b, v, j+ and j−) are the complementary properties of the solvent sys-tem. Solute and solvent properties have been described in detail in earlier publications [13–21], there are several reviews on our method [22–24], and a number of applications [25–27] and so only an outline will be given here. E is the solute excess molar refrac-tion in cm3·mol−1/10, S is a measure of the solute polarity/polarizability, A and B refer to the hydrogen bond acidity and hydrogen bond basicity of the solute, and V is McGowan’s characteristic molecular volume of the solute in cm3·mol−1/100 [28]. J+ and J− are proper-ties of positively charged and negatively charged solutes. Note that for positively charged solutes J− = 0, for negatively charges solutes J+ = 0, and for nonelectrolytes J+ =J− = 0. For the specific case of zwitterionic aminoacid solutes both J+ and J− are necessary to characterize the solutes [29].

(1)log10 (Do) = c + e ⋅ E + s ⋅ S + a ⋅ A + b ⋅ B + v ⋅ V + j+ ⋅ J

++ j− ⋅ J

750 Journal of Solution Chemistry (2019) 48:748–757

1 3

3 Results

Diffusion coefficients for ions and ionic species in organic solvents were obtained from data on single-ion conductances �o

i through the Nernst–Einstein equation, Eq. 2, where

F is the Faraday constant. We give diffusion coefficients for ions and nonelectrolytes in units of 10−5 cm2·s−1 at 298 K.

3.1 Water

Data for ions and ionic species were those that we used before [12]. For nonelectrolytes we supplemented our previous data with a considerable number of extra values of log10 (Do) [2, 5, 7, 30–34], leading to a total of 377 values, including ions, ionic species, zwitterionic amino acids and neutral nonelectrolytes. The list of solutes, diffusion coef-ficients and solute descriptors are given in Table S1 of the Supplementary Material. We found that polyaromatic hydrocarbons with three or more condensed rings were system-atically out of line, and we included an indicator variable for these solutes, Ipha, that has the value Ipha= 1 for the condensed ring compounds and Ipha= 0 for all other solutes. Our final equation is

N = 377, SD = 0.0474, R2 = 0.840, F = 241.20, PRESS = 0.91821, Q2 = 0.823, and PSD = 0.0499.

Here and elsewhere, N is the number of data points (solutes), SD is the standard devia-tion, R is the equation correlation coefficient, F is the Fisher F-statistic and PRESS and Q2 are the leave-one-out statistics. We use the predictive standard deviation, PSD, as a measure of the predictive capability of the equation [35] rather than the method of division into test and training sets, because PSD is based on the leave-one-out statistics, it takes into account the number of data points and the number of descriptors and it uses data on all the data points. The high value of the leave-one-out Q2 shows that the equation is statistically sound.

3.2 Ethanol

We obtained diffusion coefficients for ions from conductance data through Eq. 2 [36–38]. We took our previous data on nonelectrolytes [12] and added further data [39–42] to give a total of 96 values for ions plus nonelectrolytes. The resulting equation for log10 (Do) is Eq. 4. The term in Ipha was not significant. Both the value of Q2 and PSD show that the equation is statistically sound. The list of solutes, diffusion coefficients and solute descrip-tors are given in Table S2.

(2)Doi=

RT

F2�

oi

(3)log10 (D

o) = 0.1803 − 0.0119E + 0.0319 S + 0.0104A − 0.0362

− 0.2519V − 0.0148 J+ + 0.0205 J− − 0.0798 Ipha

(4)log10 (D

o) = 0.3710 − 0.1058E + 0.1369 S − 0.3189A − 0.1629

− 0.2215V + 0.4084 J+ + 0.1096 J−

751Journal of Solution Chemistry (2019) 48:748–757

1 3

N = 96, SD = 0.0666, R2 = 0.932, F = 169.17, PRESS = 0.58308, Q2 = 0.897, and PSD = 0.0823.

3.3 Methanol

There are numerous studies on single ion conductance in methanol (43–47, 48) from which we derived corresponding diffusion coefficients. Diffusion coefficients of nonelectrolytes have been well studied (4, 10, 34, 39–42, 49–53) and we collected data on a total of 96 solutes, ionic plus nonelectrolytes, that yielded Eq. 5. The list of solutes, diffusion coefficients and solute descriptors are given in Table S3.

N = 96, SD = 0.0332, R2 = 0.915, F = 135.98, PRESS = 0.1232, Q2 = 0.893, and PSD = 0.0374.

3.4 Propan‑1‑ol

Diffusion coefficients for ions were obtained from conductance data through Eq.  2 [30, 37, 43, 45]. There is surprisingly little data on diffusion coefficients of nonelectrolytes [4, 34], not enough to enable us to obtain a robust equation based on Eq. 1. However, there are considerable data on nonelectrolytes in octanol [3, 49, 50] as well as in methanol and ethanol (see above). For any given solute, a plot of log10 (Do) against the number of carbon atoms in the alcohol is almost linear. Taking into account the slight nonlinearity, it is then quite easy to interpolate values of log10 (Do) to obtain estimates in propan-1-ol (and in butan-1-ol). We obtained a total of 73 values of log10 (Do) for ions and nonelectrolytes that yielded Eq. 6. The list of solutes, diffusion coefficients and solute descriptors are given in Table S4.

N = 71, SD = 0.0487, R2 = 0.948, F = 164.81, PRESS = 0.20879, Q2 = 0.928, and PSD = 0.0576.

3.5 Butan‑1‑ol

Data are available for ions [30, 45], but, again, there is a lack of data on nonelectrolytes [4, 10, 39, 40, 42, 47]. We adopted the same stratagem as for propan-1-ol solvent and obtained data on a total of 53 solutes, leading to Eq. 7. The list of solutes, diffusion coefficients and solute descriptors are given in Table S5.

N = 53, SD = 0.0600, R2 = 0.926, F = 80.29, PRESS = 0.32266, Q2 = 0.852, and PSD = 0.0846.

(5)log10 (D

o) = 0.4601 + 0.0197E − 0.0020 S − 0.2403A − 0.0463

− 0.1377V + 0.0252 J+ − 0.0788 J−

(6)log10 (D

o) = 0.1263 − 0.0353E − 0.0144 S − 0.4433A − 0.1063

− 0.0938V − 0.1097 J+ − 0.2016 J−

(7)log10 (D

o) = 0.0608 − 0.0036E − 0.0316 S − 0.4993A − 0.0461

− 0.1822V − 0.1777 J+ − 0.2422 J−

752 Journal of Solution Chemistry (2019) 48:748–757

1 3

3.6 Octan‑1‑ol

Diffusion coefficients for ions are available from conductance data [46, 54] and there is just enough data on diffusion coefficients for nonelectrolytes [3, 10, 54–57] to be able to construct Eq. 8 for a total of 61 solutes. The list of solutes, diffusion coefficients and solute descriptors are given in Table S6.

N = 61, SD = 0.0684, R2 = 0.949, F = 141.41, PRESS = 0.3327, Q2 = 0.932, and PSD = 0.0792.

3.7 Propanone

There is a reasonable amount of data on the conductance of ions in propanone [46, 58], and several studies on the diffusion of nonelectrolytes [10, 50, 59–63]. We did not use the results of Hejtmanek and Schnelder [63] on crown ethers that were completely out of line. Interestingly, these experimental results were not well reproduced by any of the predictive methods used [63]. We compiled a total of 86 values of log10 (Do) including ions and none-lectrolytes, that yielded Eq. 9. The list of solutes, diffusion coefficients and solute descrip-tors are given in Table S7.

N = 86, SD = 0.0366, R2 = 0.837, F = 57.37, PRESS = 0.14773, Q2 = 0.767, and PSD = 0.0435.

3.8 Acetonitrile

There have been a several studies on conductance of ions in acetonitrile solvent [44, 46, 58, 64–67] so that we were able to obtain log10 (Do) for 18 ions. There are also a considerable number of studies on nonelectrolytes [10, 40, 68–74] so that we had a combined total of 74 log10 (Do) values for ions and nonelectrolytes. This led to Eq. 10. The list of solutes, diffu-sion coefficients and solute descriptors are given in Table S8.

N = 74, SD = 0.0438, R2 = 0.886, F = 73.53, PRESS = 0.15983, Q2 = 0.857, and PSD = 0.0492.

4 Discussion

There are several more solvents than those we have studied for which conductance data exists for a sufficient number of ions to be able to use in Eq. 1. Unfortunately there are not enough data on nonelectrolytes in these solvents to obtain statistically meaningful

(8)log10 (D

o) = − 0.1233 − 0.0382E − 0.0440 S − 0.5710A − 0.2862

− 0.2378V − 0.8020 J+ − 0.2899 J−

(9)log10 (D

o) = 0.6593 − 0.0221E − 0.0162 S − 0.1718A − 0.0031

− 0.1102V − 0.0167 J+ − 0.0631 J−

(10)log10 (D

o) = 0.6612 − 0.0283E − 0.0126 S − 0.2311A − 0.0511

− 0.1399V + 0.034 J+ − 0.0287 J−

753Journal of Solution Chemistry (2019) 48:748–757

1 3

equations. On the other hand, there are a large number of nonpolar solvents for which log10 (Do) are available, but for which there is no data, or very little data, on log10 (Do) for ions. However, we have results for enough solvents to show that Eq. 1 can be regarded as a general equation for the correlation and prediction of diffusion coefficients of ions and nonelectrolytes in water and in organic solvents. We include diffusion coefficients of zwitterionic aminoacids (in water), of the tetraalkyltins (in most of the solvents) and of the inorganic complexes Fe(acac)3 and Co(acac)3 as well as a number of ferrocenes (in acetonitrile) so that the scope of our equations is very large indeed. The generality of our equations is not limited by functional groups, but depends on the range of the independent variables in the various equations. We give in Table 1 the maximum and minimum val-ues of the independent variables for the nonelectrolytes studied with water and butan-1-ol solvents; that is, the solvents with the largest and smallest numbers of the diffusion coef-ficients. Even for butan-1-ol, the range of the independent variables for nonelectrolytes is large enough to include a very wide range of solute types. The statistics of our equations for log10 (Do) are all quite good, with SD values ranging from 0.0332 (methanol) to 0.0684 (octan-1-ol). By comparison, our present equation for solvent water has SD = 0.0474 for 377 ions plus nonelectrolytes as compared to our previous statistic [12] of SD = 0.085 for 190 ions plus nonelectrolytes in water.

We have data on diffusion coefficients for the large ions Ph4As+, Ph4P+ and Ph4B− but in several solvents the observed values of log10 (Do) were considerably out of line and we then omitted these ions from all of our analyses. We have no explanation for this observa-tion. Our descriptors for these ions have been used on several occasions in partition sys-tems with good results [24] so errors in the descriptors can be ruled out. Our calculated log10 (Do) values for Ph4As+and Ph4P+ were always quite close to the calculated value for Ph4B− as is required by the TATB convention [75, 76] that we used to obtain data on single

Table 1 Minimum and maximum values of the independent variables used in Eq. 1 for nonelectrolytes

Descriptor Solvent butan-1-ol Solvent water

Minimum Maximum Minimum Maximum

E 0.00 1.67 − 0.31 3.03S 0.00 2.10 − 0.21 2.78A 0.00 1.09 0.00 3.00B 0.00 2.73 0.00 3.05V 0.25 3.63 0.27 2.73

Table 2 Summary of coefficients in Eq. 1

Solvent c e s a b v j+ j−

Water 0.1752 − 0.0066 0.0373 0.0015 − 0.0348 − 0.2526 − 0.0112 0.0109Methanol 0.4601 0.0197 − 0.0020 − 0.2403 − 0.0463 − 0.1377 0.0252 − 0.0788Ethanol 0.3710 − 0.1058 0.1369 − 0.3189 − 0.1629 − 0.2215 0.4084 0.1096Propan-1-ol 0.1263 − 0.0353 − 0.0144 − 0.4433 − 0.1063 − 0.0938 − 0.1097 − 0.2016Butan-1-ol 0.0608 − 0.0036 − 0.0316 − 0.4993 − 0.0461 − 0.1822 − 0.1777 − 0.2422Octan-1-ol − 0.1233 − 0.0382 − 0.0440 − 0.5710 − 0.2862 − 0.2378 − 0.8020 − 0.2899Propanone 0.6593 − 0.0221 − 0.0162 − 0.1718 − 0.0031 − 0.1102 − 0.0167 − 0.0631Acetonitrile 0.6612 − 0.0283 − 0.0126 − 0.2311 − 0.0511 − 0.1399 0.0340 − 0.0287

754 Journal of Solution Chemistry (2019) 48:748–757

1 3

ions [24], so that the log10 (Do) values we use for single ions are compatible with the TATB convention that we use to obtain descriptors for single ions.

It is useful to compare the coefficients in Eq. 1 for the various solvents that we have studied, and these are collected in Table 2. The three most important terms for both ions and nonelectrolytes are a·A, b·B and v·V with the a-, b- and v-coefficients all being nega-tive. Thus increase in solute hydrogen bond acidity, hydrogen bond basicity and volume all lead to a decrease in the limiting diffusion coefficient, the effect of solute volume on diffu-sion being well known [77, 78]. The decrease in diffusion due to increase in solute hydro-gen bond acidity and hydrogen bond basicity can probably be attributed to increases in solute–solvent acid–base interactions. It would be useful if the coefficients in Table 2 could be used to predict coefficients for other solvents. We examined the coefficients for water and the alcohols as a function of the number of carbon atoms in the solvents. Although there are some general trends, we were unable to find statistically satisfactory correlations between any of the coefficients and the number of carbon atoms. It can be seen, almost by inspection, that the coefficients in water and in ethanol are often outliers to the general trends.

5 Conclusions

We have now shown that it is possible to correlate limiting diffusion coefficients for non-electrolytes, ions and ionic species in water and in seven organic solvents with Abraham descriptors through the linear free energy equation, LFER, Eq. 1. We can therefore now regard Eq. 1 as a general equation for the correlation of limiting diffusion coefficients. In principle, it could be applied to any solvent for which diffusion data exists on non-electro-lytes, ions and ionic species. Of course, this excludes non-polar solvents where diffusion data are restricted to non-electrolytes, but there are numerous polar or semi-polar solvents to which Eq. 1 could apply. We have not been able to extend our studies to additional sol-vents simply because of the lack of data on either non-electrolytes or ionic species. Con-ductance data exists for ions in 2-methoxyethanol [79, 80], in N,N-dimethylformamide [66, 81] and in formic acid [82], and these could be used to obtain corresponding limiting dif-fusion coefficients through Eq. 2. However, there is not enough data on non-electrolytes in these solvents to apply Eq. 1. Whether Eq. 1 could be applied to diffusion of non-electro-lytes and ions through biological or other membranes [3] remains to be seen.

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 Interna-tional License (http://creat iveco mmons .org/licen ses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

References

1. Lusis, M.A.: Predicting liquid diffusion coefficients. Chem. Process Eng. 5, 27–31 (1971) 2. Oelkers, E.H.: Calculation of diffusion coefficients for aqueous organic species at temperatures from 0

to 350 °C. Geochim. Cosmochim. Acta 55, 3515–3529 (1991) 3. Miller, D.M.: Evidence that interfacial transport is rate-limiting during passive cell membrane permea-

tion. Biochim. Biophys. Acta 1065, 75–81 (1991) 4. Lusis, M.A., Ratcliff, G.A.: Diffusion of inert and hydrogen-bonding solutes in aliphatic alcohols.

AIChE J. 17, 1492–1496 (1971)

755Journal of Solution Chemistry (2019) 48:748–757

1 3

5. Gharagheizi, F., Sattari, M.: Estimation of molecular diffusivity of pure chemicals in water: a quantita-tive structure-property relationship study. SAR and QSAR. Environ. Res. 20, 267–285 (2009)

6. Tetko, I.V., Gasteiger, J., Todeschini, R., Mauri, A., Livingstone, D., Ertl, P., Palyulin, V.A., Rad-chenko, E.V., Zefirov, N.S., Makarenko, A.S., Tanchuk, V.Y., Prokopenko, V.V.: Virtual computational chemistry laboratory—design and description. J. Comput. Aid. Mol. Des. 19, 453–463 (2005)

7. Gharagheizi, F., Eslamimanesh, A., Mohammadi, A.H., Richon, D.: Representation and prediction of molecular diffusivity of nonelectrolyte organic compounds in water at infinite dilution using the artifi-cial neural network-group contribution method. J. Chem. Eng. Data 56, 1741–1750 (2011)

8. Gharagheizi, F.: Detemination of diffusion coefficient of organic compounds in water using a simple molecular-based model. Ind. Eng. Chem. Res. 51, 2797–2808 (2012)

9. Magalhaes, A.L., Da Silva, F.A., Silva, C.M.: Tracer diffusion coefficients of polar systems. Chem. Eng. Sci. 73, 151–168 (2012)

10. Li, Z., Zuo, L., Wu, W., Chen, L.: An alternative method for correlation and evaluation of mutual dif-fusion coefficients in organic solvents at infinite dilution. J. Solution Chem. 47, 1479–1491 (2018)

11. Abraham, M.H.: Scales of solute hydrogen-bonding: their construction and application to physico-chemical and biochemical processes. Chem. Soc. Revs. 22, 73–83 (1993)

12. Hills, E.E., Abraham, M.H., Hersey, A., Bevan, C.D.: Diffusion coefficients in ethanol and in water at 298 K: linear free energy relationships. Fluid Phase Equilib. 303, 45–55 (2011)

13. Abraham, M.H., Smith, R.E., Luchtefeld, R., Boorem, A.J., Luo, R., Acree Jr., W.E.: Prediction of solubility of drugs and other compounds in organic solvents. J. Pharm. Sci. 99, 1500–1515 (2010)

14. Sedov, I.A., Magsumov, T.I., Hart, E., Higgins, E., Grover, D., Zettl, H., Zad, M., Acree Jr., W.E., Abraham, M.H.: Abraham model expressions for describing water-to-diethylene glycol and gas-to-diethylene glycol solute transfer processes at 298.15 K. J. Solution Chem. 46, 331–351 (2017)

15. Sedov, I.A., Salikov, T., Hart, E., Higgins, E., Acree Jr., W.E., Abraham, M.H.: Abraham model linear free energy relationships for describing the partitioning and solubility behavior of nonelec-trolyte organic solutes dissolved in pyridine at 298.15 K. Fluid Phase Equilib. 431, 66–74 (2017)

16. Sedov, I.A., Khaibrakhmanova, D., Hart, E., Grover, D., Zettl, H., Koshevarova, V., Dai, C., Zhang, S., Schmidt, A., Acree Jr., W.E., Abraham, M.H.: Development of Abraham model correlations for solute transfer into both 2-propoxyethanol and 2-isopropoxyethanol at 298.15 K. J. Mol. Liq. 212, 833–840 (2015)

17. Hart, E., Grover, D., Zettl, H., Koshevarova, V., Zhang, S., Dai, C., Acree Jr., W.E., Sedov, I.A., Stolov, M.A., Abraham, M.H.: Abraham model correlations for solute transfer into 2-methoxyetha-nol from water and from the gas phase. J. Mol. Liq. 209, 738–744 (2015)

18. Sedov, I.A., Stolov, M.A., Hart, E., Grover, D., Zettl, H., Koshevarova, V., Dai, C., Zhang, S., Acree Jr., W.E., Abraham, M.H.: Abraham model correlations for describing solute transfer into 2-butoxyethanol from both water and the gas phase at 298 K. J. Mol. Liq. 209, 196–202 (2015)

19. Abraham, M.H., Acree Jr., W.E.: Equations for the partition of neutral molecules, ions and ionic species from water to water–methanol mixtures. J. Solution Chem. 45, 861–874 (2016)

20. Abraham, M.H., Acree Jr., W.E.: Partition coefficients and solubilities of compounds in the water–ethanol solvent system. J. Solution Chem. 40, 1279–1290 (2011)

21. Abraham, M.H., Acree Jr., W.E.: Equations for the partition of neutral molecules, ions and ionic species from water to water–ethanol mixtures. J. Solution Chem. 41, 730–740 (2012)

22. Clarke, E.D., Mallon, L.: The determination of Abraham descriptors and their application to crop protection research. In: Jeschke, P., Kramer, W., Schirmer, U., Witschel, M. (eds.) Modern Methods in Crop Protection Research. Wiley, Hoboken (2012)

23. Endo, S., Goss, K.-U.: Applications of polyparameter linear free energy relationships in environ-mental chemistry. Environ. Sci. Technol. 48, 12477–12491 (2014)

24. Abraham, M.H., Acree Jr., W.E.: Descriptors for ions and ion-pairs for use in linear free energy relationships. J. Chromatogr. A 1430, 2–14 (2016)

25. Abraham, M.H.: Human intestinal absorption—neutral molecules and ionic species. J. Pharm. Sci. 103, 1956–1966 (2014)

26. Zhang, K., Abraham, M.H., Liu, X.: An equation for the prediction of human skin permeability of neutral molecules, ions and ionic species. Int. J. Pharmaceutics 521, 259–266 (2017)

27. Zhang, K., Fahr, A., Abraham, M.H., Acree Jr., W.E., Tobin, D.J., Liu, X.: Comparison of lipid membrane–water partitioning with various organic solvent–water partitions of neutral species and ionic species: uniqueness of cerasome as a model for the stratum corneum in partition processes. Int. J. Pharmaceutics 494, 1–8 (2015)

28. Abraham, M.H., McGowan, J.C.: The use of characteristic volumes to measure cavity terms in reversed-phase liquid chromatography. Chromatographia 23, 243–246 (1987)

756 Journal of Solution Chemistry (2019) 48:748–757

1 3

29. Abraham, M.H., Acree Jr., W.E.: Solvation descriptors for zwitterionic α-aminoacids; estimation of water–solvent partition coefficients, solubilities and hydrogen-bond acidity and hydrogen-bond basicity. ACS Omega (2019). https ://doi.org/10.1021/acsom ega.8b032 42

30. Zhao, Y., Freeman, G.R.: Electrical conductance of tetraphenylphosphonium and tetraphenylboride salts in C1 to C4 alcohols. J. Phys. Chem. 100, 17568–17572 (1996)

31. Umecky, T., Kuga, T., Funazukuri, T.: Infinite dilution binary diffusion coefficients of several amino acids in water over a temperature range from (293.2 to 333.2) K with the Taylor dispersion technique. J. Chem. Eng. Data 51, 1705–1710 (2006)

32. Longsworth, L.G.: Diffusion measurements, at 25 °C, of aqueous solutions of amino acids, peptides and sugars. J. Am. Chem. Soc. 75, 5705–5709 (1953)

33. Mogi, N., Sugai, E., Fuse, Y., Funazukuri, T.: Infinite binary diffusion coefficients for six sugars at 0.1 MPa and temperatures from (273.2 to 353.2) K. J. Chem. Eng. Data 52, 40–43 (2007)

34. Arai, C., Hozumi, K., Sano, Y., Matsuda, H.: Diffusion coefficients of aromatic compounds in water and alcohols. Kagaku Kogaku Ronbunshu 18, 538–541 (1992)

35. Abraham, M.H., Acree Jr., W.E., Leo, A.J., Hoekman, D.: The partition of compounds from water and from air into wet and dry ketones. New J. Chem. 33, 568–573 (2009)

36. Barthel, J., Neueder, R., Feuerlein, F., Strasser, F., Iberi, L.: Conductance of electrolytes in ethanol solutions from − 45 to 25 °C. J. Solution Chem. 12, 449–471 (1983)

37. Evans, D.F., Gardam, P.: Transport properties in hydrogen-bonding solvents. I. The conductance of teraalkylammonium salts in ethanol and propanol at 25 °C. J. Phys. Chem. 72, 3281–3286 (1968)

38. Kay, R.L., Evans, D.F.: The effect of structure on the mobility of symmetrical ions in aqueous solu-tion. J. Phys. Chem. 70, 2325–2335 (1966)

39. Chen, H.-C., Chen, S.-H.: Diffusion of crown ethers in alcohols. J. Phys. Chem. 88, 5118–5121 (1984) 40. Tominaga, T., Tenma, S., Watanabe, H.: Diffusion of cyclohexane and cyclopentane derivatives in

some polar and non-polar solvents. J. Chem. Soc. Faraday Trans. 92, 1863–1867 (1996) 41. Chan, T.C., Chen, N., Lu, J.G.: Diffusion of disubstituted aromatic compounds in ethanol. J. Phys.

Chem. A 102, 9087–9090 (1998) 42. Evans, D.F., Tominaga, T., Chan, C.: Diffusion of symmetrical and spherical solutes in protic, aprotic

and hydrocarbon solvents. J. Solution Chem. 8, 461–478 (1979) 43. Kay, R.L., Zawoyski, C., Evans, D.F.: The conductance of the symmetrical tetralkylammonium halides

and picrates in methanol at 25 and 10°. J. Phys. Chem. 69, 4208–4215 (1965) 44. Kay, R.L., Evans, D.F., Cunningham, G.P.: The conductance of large hydrophobic ions in water, meth-

anol and acetonitrile at 25 °C. J. Phys. Chem. 73, 3322–3326 (1969) 45. Barthel, J., Krell, M., Iberl, L., Feuerlein, F.: Conductance of 1:1 electrolytes in methanol solutions

from − 45 to + 25 °C. J. Electroanal. Chem. 214, 485–505 (1986) 46. Krumgalz, B.S.: Separation of limiting equivalent conductances into ionic contributions in non-aque-

ous solutions by indirect methods. J. Chem. Soc. Faraday Trans. I 79, 571–587 (1983) 47. Wawrzyniak, G., Jasinski, T.: Solvation of electrolytes in a binary mixture of solvents propylene car-

bonate–methanol. Polish J. Chem. 59, 1181–1192 (1985) 48. Chowdhuri, S., Chandra, A.: Solute size effects on the solvation structure and diffusion of ions in

methanol. J. Chem. Phys. 124, 084507 (2006) 49. Chan, T.C.: Diffusion of pseudospherical molecules: an investigation on the effects of dipole moment.

J. Chem. Phys. 79, 3591–3593 (1983) 50. Chan, M.L., Chan, T.C.: Diffusion of pseudoplanar solutes: an investigation on the effects of hydrogen

bonding. J. Phys. Chem. 99, 5765–5768 (1995) 51. Longsworth, L.G.: Diffusion in the water–methanol system and the Walden product. J. Phys. Chem.

67, 689–693 (1963) 52. Safi, A., Nicolas, C., Neau, E., Chevalier, J.-L.: Diffusion coefficients of organic compounds at infinite

dilution in mixtures involving associating compounds. Experimental determination and modeling by group contribution methods. J. Chem. Eng. Data 53, 444–448 (2008)

53. Chan, T.C., Tang, W.K.: Diffusion of aromatic compounds in nonaqueous solvents: a study of solute, solvent and temperature dependence. J. Chem. Phys. 138, 224503 (2013)

54. Chen, R., Freeman, G.R.: Solvent effects on reactivity of solvated electrons with nitrate ion and on electrolyte conductivity in C1 to C10 n-alcohols. J. Phys. Chem. 99, 4970–4975 (1995)

55. Chen, S.-H., Evans, D.F., Davis, H.T.: Tracer diffusion in methanol, 1-butanol and 1-octanol from 298 to 433 K. AIChE J. 29, 640–645 (1983)

56. Erkey, C., Alhamid, K.A., Akgerman, A.: Investigation of the effects of molecular association on diffu-sion in binary liquid mixtures at the infinite dilution limit. J. Chem. Phys. 94, 3867–3871 (1991)

57. Gabler, T., Paschke, A., Schuurmann, G.: Diffusion coefficients of substituted benzenes and alcohols at high dilution in octan-1-ol. J. Chem. Eng. Data 43, 413–416 (1998)

757Journal of Solution Chemistry (2019) 48:748–757

1 3

58. Evans, D.F., Zawoyski, C., Kay, R.L.: The conductance of the symmetrical tetraalkylammonium hal-ides and picrates in acetonitrile at 25 °C. J. Phys. Chem. 69, 3878–3885 (1965)

59. Chan, T.C., Ma, N.L., Chen, N.: The effects of molecular association on mutual diffusion in acetone. J. Chem. Phys. 107, 1890–1895 (1997)

60. Chen, S., Xu, J., Chan, T.C.: Steric effects on diffusion of associated molecules in acetone. Chem. Commun. 8, 898–899 (2002)

61. Chan, T.C.: Diffusion of aromatic compounds: an investigation on the effects of molecular shape, mass and dipole moment. J. Chem. Phys. 80, 5862–5864 (1984)

62. Chan, T.C., Chan, M.L.: Diffusion of pseudo-planar molecules: an experimental evaluation of the molecular effects on diffusion. J. Chem. Soc. Faraday Trans. 88, 2371–2374 (1992)

63. Hejtmanek, V., Schnelder, P.: Diffusion coefficients of some large organic molecules in acetone, cyclohexane and toluene. J. Chem. Eng. Data 38, 407–409 (1993)

64. Singh, P., MacLeod, I.D., Parker, A.J.: Solvation of ions. Part XXIX. Ionic conductances in acetoni-trile–water and pyridine–water mixtures. J. Solution Chem. 13, 103–119 (1984)

65. Safonova, L.P., Patsatsiya, B.K., Kolker, A.M.: Effect of temperature on the electric conductivity of individual ions in acetonitrile. Zh. Fiz. Khim. 68, 44–47 (1994)

66. Bhattacharya, S.K., Ukil, U., Kundu, K.K.: Ion conductance and studies of KI and some reference electrolytes KKBPh4, Ph4AsI and n-Bu4NI in acetonitrile, N, N-dimethylformamide and their isodi-electric binary mixtures at 298.15 K. Indian J. Chem. 35A, 1038–1046 (1996)

67. Porras, S.P., Kenndler, E.: Formamide as solvent for capillary zone electrophoresis. Electrophoresis 25, 2946–2968 (2004)

68. Easteal, A.J., Woolf, L.A.: Solute–solvent interaction effects on tracer diffusion coefficients. J. Chem. Soc. Faraday Trans. 80, 1287–1295 (1984)

69. Valencia, D.P., Gonzalez, F.J.: Estimation of diffusion coefficients by using a linear correlation between the diffusion coefficient and molecular weight. J. Electroanal. Chem. 681, 121–126 (2012)

70. Saiki, H., Takami, K.: Tominaga, T,: diffusion of porphyrins and quinones in organic solvents. Phys. Chem. Chem. Phys. 1, 303–306 (1999)

71. Toriumi, M., Katooka, R., Yui, K., Funazukuri, T., Kong, C.Y., Kagei, S.: Measurement of binary diffusion coefficients for metal complexes in organic solvents by the Taylor dispersion method. Fluid Phase Equil. 297, 62–66 (2010)

72. Ikeychi, H., Naganuma, K., Ichikawa, M., Ozawa, H., Ino, T., Sato, M., Yonezawa, H., Mukaida, S., Yamamoto, A., Hashimoto, T.: Diffusion coefficients of tris(-diketonato)ruthenium complexes of dif-ferent charge numbers in acetonitrile solutions, measured by chronoamperometry. J. Solution Chem. 36, 1243–1259 (2007)

73. Li, J., Carr, P.W.: Accuracy of empirical correlations for estimating diffusion coefficients in aqueous organic mixtures. Anal. Chem. 69, 2530–2536 (1997)

74. Ye, F., Jensen, H., Larsen, S.W., Yaghmur, A., Larsen, C., Ostergaard, J.: Measurement of drug dif-fusivities in pharmaceutical solvents using Taylor dispersion analysis. J. Pharm. Biomed. Anal. 61, 176–183 (2012)

75. Kalidas, C., Hefter, G., Marcus, Y.: Gibbs energies of transfer of cations from water to mixed aqueous organic solvents. Chem. Rev. 100, 819–852 (2002)

76. Marcus, Y.: Gibbs energies of transfer of anions from water to mixed aqueous organic solvents. Chem. Rev. 107, 3880–3897 (2007)

77. Othmar, D.F., Thakar, M.S.: Correlating diffusion coefficients in liquids. Ind. Eng. Chem. 45, 589–593 (1953)

78. Wilke, C.R., Chang, P.: Correlation of diffusion coefficients in dilute solutions. AIChE J. 1, 264–270 (1955) 79. Victor, P.J., Muhuri, P.K., Das, B., Hazra, D.K.: Thermodynamics of ion association and solvation in

2-methoxyethanol: behavior of tetraphenylarsonium, picrate and tetraphenylborate ions from conduc-tivity and ultrasonic data. J. Phys. Chem. B 103, 11227–11232 (1999)

80. Victor, P.J., Muhuri, P.K., Das, B., Hazra, D.K.: Thermodynamics of ion association of tetraphe-nylphosphonium, tetraphenylarsonium, and some common cations in 2-methoxyethanol using conduc-tometry and FT-Raman spectroscopy. J. Phys. Chem. B 104, 5250–5356 (2000)

81. Safonova, L.P., Sakharov, D.V., Shmukler, L.E., Kolker, A.M.: Conductance studies of 1:1 electrolytes in N, N-dimethylformamide at various temperatures. Phys. Chem. Chem. Phys. 3, 819–823 (2001)

82. Paul, R.C., Sharma, R., Puri, T., Kapoor, R.: Formic acid solvent system. II. Conductance and solva-tion studies of 1:1 electrolytes in formic acid. Aust. J. Chem. 30, 535–5471 (1977)

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.